You are on page 1of 19

Clay Minerals (1999) 34, 735

The origin and formation of clay minerals


in soils" past, present and future
perspectives
M . J. W I L S O N
Macaulay Land Use Research Institute, Craigiebuckler, Aberdeen AB15 8QH, UK
(Received 23 September 1997," revised 15 January 1998)

A B S T R A C T : The origin and formation of soil clay minerals, namely micas, vermiculites,
smectites, chlorites and interlayered minerals, interstratified minerals and kaolin minerals, are
broadly reviewed in the context of research over the past half century. In particular, the pioneer
overviews of Millot, Pedro and Duchaufour in France and of Jackson in the USA, are considered in
the light of selected examples from the huge volume of work that has since taken place on this topic.
It is concluded that these early overviews may still be regarded as being generally valid, although it
may be that too much emphasis has been placed upon transformation mechanisms and not enough
upon neoformation processes. This review also highlights some of the many problems pertaining to
the origin and formation of soil clays that remain to be resolved.

It has long been recognized that the minerals in the


clay (<2 gm) fractions of soils play a crucial role in
determining their major physical and chemical
properties, and inevitably, questions concerning
the origin and formation of these minerals have
assumed some prominence in soil science research.
This review considers some important aspects of
these questions and is confined to clay minerals in
soils as they are generally understood, that is as a
medium for plant growth.
The occurrence of clay minerals in saprolites or
weathered rock is not discussed to any great extent
because many studies have shown that saprolites
are characterized by physicochemical conditions
that are different to those in soils. Intimate grain-tograin contacts promote a special chemical environment on a local scale, bringing about the formation
of transient clay mineral phases which quickly
disappear in the overlying soil.
For the purposes of this review the perspective of
the past will be taken to be that of the early 1960s,
which is a convenient historical benchmark in that
it represents a time when the overall structural
characteristics of the layer silicate clay minerals
were to a large extent known, even though much

detail remained to be filled in, as well as a time that


immediately pre-dated the widespread utilization in
soil science of analytical techniques such as
scanning electron microscopy, electron probe
microanalysis, M6ssbauer spectroscopy, electron
spin resonance spectroscopy and infrared spectroscopy. This review will therefore attempt to
summarize the general conclusions that had been
arrived at regarding the origin and formation of
certain clay minerals in soils in the early 1960s, the
developments since that time and the situation as it
currently stands, and to consider the outstanding
problems to be addressed in the future. In this
context, particular attention will be paid to micas,
vermiculites, smectites, chlorites, intcrstratified
minerals and kaolin minerals. As in all such
reviews, a certain amount of subjectivity and
personal preference is perhaps inevitable.
PAST

PERSPECTIVES

The reviews of Millot (1965) and Jackson (1964)


represent a convenient starting point with regard to
the origin and formation of soil clays as assessed
principally by the X-ray diffraction and microscopic

9 1999 The Mineralogical Society

M.J. Wilson

techniques that were available at the time. Millot


distinguished three principal processes to account
for the genesis of clay minerals, which may occur
at different points in the geochemical cycle
including weathering or soil formation at the
earth's surface. These processes are: (a) detrital
inheritance whereby, for soils, clay minerals are
inherited from pre-existing parent rock or weathered
materials; (b) transformation where the essential
silicate structure of the clay mineral is maintained
to a large extent, but with major change in the
interlayer region of the structure; and (c) neoformation, where the clay mineral forms through crystallization of gels or solutions.
Inherited soil clays may be of an extremely
diverse and complex nature, reflecting both the
variety of the parent rock as well as the
transformation and neoformation processes that
may have occurred in previous weathering environments. In order to understand current clay mineralforming processes in a soil, it is essential that the
contribution of inheritance is clearly understood.
With regard to transformation of clay minerals, a
general example would be:
illite ~ vermiculite ~ smectite.
This reaction proceeds through a process of
depletion and exchange of interlayer K and
concomitant decrease of layer charge. Such
changes are, however, deceptively simple and
have given rise to much debate about the precise
mechanisms involved, as will be discussed later.
Millot distinguished 'degradation' and 'aggradation'
as separate forms of the transformation process. The
above conversion of illite to montmorillonite
involves depletion of elements from illite and is
termed degradation, but the reverse reaction
(aggradation) involves addition of K and other
elements. Millot considered that degradation was
characteristic of weathering rocks and soils, but that
aggradation was rare in such environments.
The formation of clay minerals through neoformation clearly depends upon the appropriate
physicochemical conditions of the immediate
weathering environment, such as the pH, composition and concentration of the soil solutions, as well
as the nature of the starting material and factors
relating to the external environment like temperature, rainfall and percolation rate. Millot described
kaolinite and montmorillonite as classical products
of neoformation in soils forming under contrasting
conditions. Thus, kaolinite is typical of freely

drained, acid and base-depleted tropical environments, where an abundant supply of water ensures
the required silica and alumina. Montmorillonite is
more typical of a poorly drained or hydromor-phic
soils under alkaline conditions, rich in Mg and Ca
ions and where Si, A1 and commonly Fe tend to
accumulate.
These generalizations were further developed and
synthesized by French soil scientists, most notably
Pedro and Duchaufour. Pedro (1964) distinguished
two mineral weathering processes implicated in the
formation of clay minerals in soils which he termed
'hydrolysis' and 'acidolysis'. (A summary of this
work in English is given in Pedro (1982)).
Hydrolysis of minerals occurs through dilute
solutions in the pH range 5-9.6 and may be total
or partial. Total hydrolysis leads to the removal of
all elements including silica, and to the precipitation of gibbsite and kaolinite minerals, whereas
partial hydrolysis, under different conditions, leads
to the formation of smectite minerals (Pedro, 1982).
Acidolysis operates when the soil solution pH is <5
or has strongly complexing properties and can again
be total or limited. Total acidolysis involves
complete solubilization of minerals with no
precipitation of AI. Limited acidolysis leads to the
fixation of AI in octahedral and interlamellar
positions in layer silicates. In general, acidolysis
is associated with podzols, podzolic brown soils and
acid brown soils of cold temperate climates,
whereas hydrolysis is dominant in ferrallitic soils
of the warm humid tropics and is prominent in
warm temperate zones and in the dry subtropics
(Table 1).
A similar synthesis relating to soil clays and
weathering was described by Duchaufour (1960)
who distinguished geochemical weathering under
near neutral conditions with no organic acid anions
and typical of tropical environments, and biochemical weathering under acid conditions with organic
anions and typical of temperate climates. The
former is characterized by neoformed clay minerals
and the latter by clay minerals formed by
transformation.
The overview of Jackson (1964) concerning the
distribution, stability and weathering reactions of
clay minerals in soils is, largely, consistent with the
conclusions of the French soil scientists. Jackson
described the dominant clay mineral types occurring in the Soil Orders of the taxonomy of the
USDA Soil Survey Staff (1960). There was a
predominance of micaceous, interstratified or

Origin and formation of clay minerals in soils


TABLE l. Occurrence of soil clay minerals in relation to weathering process, principal
mechanism and soil types (Adapted from Pedro, 1982).
Weathering
process

Principal
mechanism

Transformation

Acidolysis
(partial)

Neoformation

Hydrolysis
(partial)

Clay
minerals

Hydrolysis
(total)

interlayered layer silicate minerals in soils such as


Entisols, Inceptisols and Spodosols which occur
widely in temperate climates. Clay minerals in soils
of tropical climates such as Oxisols and Ultisols,
were dominated by kaolinite and halloysite, in
addition to gibbsite and sesquioxides. However, the
occurrence of 2:1 to 2:2 intergrades was also
indicated in both Oxisols and Ultisols and in the
latter, vermiculite was also mentioned. In the
weathering reactions described by Jackson the
emphasis was placed very much upon transformation pathways (Fig. 1) and particularly on the role
of hydroxy A1 interlayers. Thus, it was envisaged
that both smectite and vermiculite minerals could
accumulate aluminium interlayers during extensive
weathering, to the point where they approach a
mineral resembling dioctahedral chlorite. A subsequent weathering step involving tetrahedra inversion could then give kaolinite (or halloysite),
although kaolinite could also form from mica
directly. Kaolinite formation was thus proposed as
proceeding through 2 : 1 ~ 2 : 2 ~ 1 : 1 or through

MICAS

CHLORITE

~'~

VERMICULITE

VERMICULITE

~
/~

PEDOGENIC
INTERGRADES

Principal
soil types

Smectites
Vermiculites
Al-intergrades
Al-chlorite
Smectites

Spodosols
Inceptisols
Entisols

Kaolinite
Halloysite
Gibbsite

Ultisols
Oxisols

Spodosols
Mollisols
Alfisols

2:1--* 1:1 pedogeochemical reactions. A major role


for precursor crystals was also suggested in the
growth of montmorillonite with hydroxy units of
A1, Fe and Mg forming on pre-existing montmorillonite crystallites, which become silicated to
form new montmorillonite layers. Jackson (1964)
concluded that "the almost universal occurrence of
hydroxy cation units precipitated on the surfaces of
layer silicates exposed to weathering is thus
extremely fundamental to clay mineralogy and
clay genesis".
PRESENT

PERSPECTIVES

This section is concerned with research bearing on


the origin of clay minerals in soils, in a global
context, which post-dates the seminal French and
American work described above. In particular, it is
of interest to consider whether the major conclusions reached need to be modified, refined or even
abandoned in the light of the enormously increased
volume of data now available on clay minerals in

MONTMORILLONITE

~
/~

AL
CHLORITE

/~

PEDOGENIC
INTERGRADES

KAOLINITE
HALLOYSITE

FIG. 1. Pathways for the formation of soil clay minerals as outlined by Jackson (1964).

GIBBSITE

10

M. J. Wilson

soils, often gained through the utilization of new


techniques. Comprehensive reviews of the literature
on soil clay minerals up until 1989 are available in
the relevant chapters of the monograph entitled
Minerals in Soil Environments (Dixon & Weed,
1989) published by the Soil Science Society of
America. This forms a very useful starting point for
the present review where the emphasis will be
placed upon recent work.
M i c a minerals'

It is now clear, as it has been for some


considerable time (Fanning et al., 1989), that the
micaceous minerals that occur in soil clay fractions
have been inherited largely from parent rock or
other material, where they originally formed under
different P - T conditions from those pertaining at
the earth's surface. However, there is some
evidence that mica may form pedogenically, but
only in special circumstances. Thus, Niederbudde
(1975) and Niederbudde & Kussmaul (1978)
presented evidence to support the formation of
mica-like clay minerals in fertilized loess-derived
soils in Germany following K fixation by beidellitic
smectites. The conversion of K-saturated smectites
to mica-like products following extensive wetting
and drying cycles in the laboratory, and involving
change from turbostratic stacking to a semi-ordered
structure (de la Calle & Suquet, 1988), suggests a
possible mechanism for such a change. Similarly, it
has been suggested that the clay micas found in the
surface horizons of soils of arid environments, such
as south west USA (Nettleton et al., 1973) and Iran
(Mahjoory, 1975) are pedogenic in origin. In both
instances a transformation process was envisaged
whereby it was possible to fix K in pre-existing
smectites because of the hot and dry soil conditions.
The need for caution in proposing mica
formation by pedogenesis is emphasized by
previously mistaken interpretations concerning
concentrations of clay mica in the surface horizons
of Hawaiian soils. Initially, it was thought that this
mica must be pedogenic because mica was
undetectable in the sub-surface soil and the basaltic
parent materials (Juang & Uehara, 1968) but later
work using isotopic analysis showed conclusively
that the mica mineral was older than the basalt flow
from which the soil was derived and therefore could
not be pedogenic (Dymond et al., 1974). Aeolian
deposition following tropospheric transport almost
certainly explains the origin of the mica minerals in

these H a w a i i a n soils. However, c o n v i n c i n g


evidence for neoformed mica in some Australian
soils was presented by Norrish & Pickering (1983).
A neoformational origin was indicated by the
perfect platy hexagonal morphology of the clay
mica and it is significant that the mineral is Fe-rich,
in this respect resembling glauconite which forms
in surface sediments under marine conditions. It is
not clear, however, that the mica described by
Norrish & Pickering (1983) is truly pedogenic in
that the soil was derived from lacustrine sediments.
It is possible that the mica may have formed in
special conditions obtained during the drying out of
the lake bed, rather than during soil formation on
the lake bed sediments.
A general point made by Fanning et al. (1989)
was that, at that time, there were very few detailed
studies of the crystallochemistry of soil clay micas.
This is true even now but there have been some
interesting advances in the nineties brought about
by the use of analytical electron microscopy,
p a r t i c u l a r l y by Robert and co-workers in
Versailles. Although it may be agreed that virtually
all soil micas are of inherited origin, it is still
important to determine what changes may have
occurred to the clay size micas under pedogenetic
conditions, bearing in mind the importance
attributed to micas as precursors of interstratified,
expansible and interlayered minerals in soils.
Robert et al. (1991) using HRTEM on impregnated
sections of French soils derived from sediments,
showed that many of the particles were exceedingly
thin, recalling the fundamental illite particles
described by Nadeau et al. (1984) in their studies
of interstratified clays. Similar findings for clay-size
micas in Iowa soils were arrived at independently
by Laird & Nater (1993). Robert et al. (1991) went
on to distinguish two types of mica mineral in the
clay fractions of the soils they studied (Table 2).
The first, described as micromica, occurs in
TABLE2. Nature of i//itic clays in soils (after Robert et
al., 1991).
Characteristics
a b plane
Thickness
K20
Charge
Structure

Micromica

Illite (sensu stricto)

micrometric
>10 nm
-10%
>0.9
3-dimensional

nanometric
~5 nm
-7.5%
<0.6
3-dimensional

Origin and Jbrmation of clay minerals' in soils

relatively thick particles of micrometric size with a


full content of K and is highly charged. The second
type is termed illite (sensu stricto) and occurs in
thin particles of nanometric size, is somewhat
depleted in K and is less highly charged. These
two types of micaceous clays can exist in composite
crystals, implying that there is a genetic relationship
between them. Later work on the weathering of
micaceous clays in soils (Romero et al., 1992;
Aouidjit et al., 1996) does indeed show that particle
microdivision is an important part of the process.
Recent studies of Spanish soils developed upon
highly micaceous parent materials presented
chemical and structural evidence for the conversion
of 2MI muscovitic mica to the 1Md polytype with
decreasing grain size (Martin Garcia et al., 1997).
The picture that is emerging from these investigations is that illitic clays may evolve in soil from the
weathering of dioctahedral muscovitic mica
minerals involving loss of K and layer charge and
principally mediated by the physical process of
particle microdivision.
Although the vast majority of micaceous soil
clays are dioctahedral and aluminous, it should not
be forgotten that trioctahedral micaceous soil clays
do exist. Thus, Fordham (1990a) found that the
weathering of granite in Australia, in conditions that
led to the formation of lateritic soils, yielded
trioctahedral micaceous clays in the initial stages
of the weathering sequence. Again, Wilson et al.
(1997) recorded trioctahedral micaceous clays
accompanied by kaolinite in some alluvial soils
from Nigeria. In both of these cases the trioctahedral clays appeared to be derived from biotite and
this raises the question as to why vermiculitization
of the biotite did not occur, as is commonly found
in soils of humid-temperate climates. It may be
relevant that in tropical soils, a direct biotite~kaolinite transformation can occur without an intervening vermiculite stage (De Kimpe & Tardy,
1968). It should also be noted that oxidized biotites
retain their K more strongly than unoxidized
biotites and are hence more resistant to vermiculitization. A possible explanation for the rather
unexpected occurrence of trioctahedral micaceous
clays in tropical soils is, therefore, that the
precursor biotites were oxidized and weathered
directly to kaolinite. Such kaolinite formation does
not take place homogeneously throughout the
biotite crystals, but tends to affect outer surfaces
and cleavage planes exposed by exfoliation
(Ojanuga, 1973) and in these circumstances it is

11

conceivable that oxidized but otherwise relatively


unaltered biotite fragments may eventually find
their way into soil clay fractions.
V e r m i c u l i t e minerals'

The review of Douglas (1989) concerning the


origin and formation of vermiculite minerals in
soils emphasized a number of significant points.
Firstly, there is general agreement that most
vermiculites are considered to form by the weathering of mica as was illustrated early on by the
work of Walker (1949). It is true that there is some
evidence to support the formation of vermiculite by
the weathering of chlorite, involving the preferential
decomposition of the interlayer hydroxide sheet
(Stephen, 1952; Ross & Kodama, 1976; Makumbi
& Herbillon, 1972), or even through the decomposition of non layer-silicate minerals such as
orthopyroxene (Basham, 1974), but in most
instances there seems no need to look further than
a genetic relationship with a precursor mica. This
link is particularly obvious in the many studies that
show vermiculite increasing towards the surface of
soil profiles with a concomitant decrease in the
intensity of the 10 A mica peak. The most detailed
studies have focused on the biotite~vermiculite
transformation.
The structure of vermiculite itself, following
principally upon the fundamental studies of de la
Calle and co-workers (references in de la Calle &
Suquet, 1988), is well understood and the structures
of both the silicate layers and the interlamellar
space, as well as the various ways in which the
layers may be stacked according to the saturating
cation and relative humidity, have been elucidated.
The vermiculitization of biotite involves a number
of factors that have been explored in laboratory
experiments. For example, the release of interlayer
K to an external solution may take place by a
diffusion process requiring that the concentration of
K in the solution is below a critical level of 11 ppm
(Newman, 1969). The vermiculitization reaction
ceases if this critical level is exceeded. Numerous
studies have shown that vermiculitization of biotite
also involves the oxidation of structural Fe (Farmer
et al., 1971; Vicente-Hernandez et al., 1983) and
that in order to maintain overall electrical neutrality
of the structure, this process results in the expulsion
of Fe from octahedral sites. It is not clear whether
oxidation of biotite can occur in soils without prior
loss of K. Laboratory experiments indicate that loss

12

M: J. Wilson

of K is necessary, but the natural occurrence of


oxybiotites as well as evidence that a direct biotite
to kaolinite weathering step is possible, as indicated
in the above discussion on trioctahedral clay micas,
suggest that these experiments may not necessarily
apply to all soil conditions.
Whether or not previous vermiculitization of
biotite is essential for oxidation of Fe to occur, the
loss of Fe effectively leads to a more dioctahedral
structure which in turn leads to a re-orientation of
octahedral hydroxyl from a direction normal to the
layer structure in unaltered biotite to a direction that
is more inclined to the layer structure in oxidized
biotite (Juo & White, 1969). This effectively places
the interlayer K into a less negative environment so
that it is consequently held more tightly within the
structure (Norrish, 1973). The changes involving
oxidation and loss of Fe have been suggested as
explanations for the loss of layer charge in the
biotite to vermiculite transformation, but other
mechanisms may also be involved such as loss of
hydroxyl ions.
Despite the weight of research effort on
trioctahedral vermiculites, it is a fact that dioctahedral vermiculite is much more common in soil
clays. This mineral, first discovered in British soils
(Brown, 1953) occurs particularly in severely
weathered soils such as Ultisols, Alfisols or even
Oxisols and is also found commonly in the upper
horizons of Spodosols (Douglas, 1989). Frequently,
dioctahedral vermiculite is interlayered with nonexchangeable hydroxy AI. The reason dioctahedral
vermiculitic minerals have not been so extensively
studied is that they are found largely in clay
fractions, where they are difficult to purify. A
muscovite to dioctahedral vermiculite transformation analogous to the biotite to trioctahedral
vermiculite transformation in the non-clay fractions
of soils has rarely, if ever, been observed, and
mucovitic mica cannot be readily vermiculitized in
the laboratory without resort to extreme chemical
techniques. Thus, there is no ready source of
material with which to study experimentally such
problems as the mechanisms involved in the loss of
layer charge in dioctahedral vermiculite as there is
for their trioctahedral counterparts.
Although the formation of dioctahedral vermiculite from a precursor dioctahedral mica seems
probable in the majority of cases, it is of interest
that other pathways have been suggested. Thus, in
an early study, Barshad & Kishk (1969) presented
evidence to support formation as precipitation

products, but this proposal does not seem to have


been widely accepted. Fordham (1990b) suggested,
on the basis of analytical electron microscopy, that
in a granitic weathering profile in Australia,
trioctahedral vermiculite converted to dioctahedral
vermiculite. It seems unlikely, however, that such a
mechanism could explain the widespread occurrence of dioctahedral vermiculitic minerals in soils.
Recent work by Aouidjit et al. (1996), however,
using high resolution transmission electron microscopy (HRTEM) presented lattice images showing
that vermiculitization of muscovitic-type mica did
occur in very fine particles engendered by
microdivision. In particular, they showed micrographs showing particles with a muscovitic core
and a vermiculitic rim consisting of ~15 layers as
well as small vermiculitic particles derived from
muscovite consisting of 2 to 5 layers in thickness.
This work suggests that very fine grained mica
particles could be more susceptible to vermiculitization in the soil by reason of their higher specific
surface and enhanced reactivity.
Smectite minerals

The factors that strongly influence the origin and


formation of smectites in soils, as reviewed by
Borchardt (1989), include low-lying topography,
poor drainage and base-rich parent material, leading
to favourable chemical conditions characterized by
high pH, high silica activity and an abundance of
basic cations. These conditions are met in many
soils under temperate or cold climates or even in
tropical climates where leaching is limited for
various reasons, including low precipitation, a
horizon in the soil profile that impedes the
passage of water or a naturally high water table
(Allen & Hajek, 1989).
Soil smectites are overwhelmingly dioctahedral,
with only montmorillonite and beidellite being of
any real importance. However, from a review of
published chemical analyses, Wilson (1987)
concluded that soil smectites were somewhat more
Fe-rich than montmorillonite and beidellite sensu
stricto and tended to show intermediate charge
characteristics with many falling into the category
of ferruginous beidellites. These observations imply
that soil smectites could be somewhat different to
those associated with bentonite and other geological
deposits, which is perhaps not surprising considering the range of chemical conditions that may
occur in soils as well as the three distinct pathways

Origin and formation of clay minerals in soils

(inheritance, neoformation and transformation) that


may be involved in soil smectite formation. Also, in
the soil environment there is often no shortage of
Fe, and pedogenic conditions often promote its
solubilization and mobilization.
Wilson (1993) reviewed the literature in order to
assess the relative importance of the three pathways
for smectite formation in the Orders of Soil
Taxonomy and some of the main points are as
follows.
In Entisols, smectites are often inherited. Thus,
inherited trioctahedral smectite is found in the
proto-soils forming from the 1980 Mount Saint
Helens pyroclastic flow (Pevear et al., 1982; La
Manna & Ugolini, 1987) and inherited smectite
may occur in Entisols under an extreme range of
climatic conditions as, for example, in the alluvial
soils of the Blue Nile Clay Plains of Sudan (Wilson
& Mitchell, 1979) and in Orthents and Psamments
on altered basalt flows in the Faeroe islands
(Rutherford & Debenham, 1981). Neoformational
smectite in Entisols may encompass a similar
climatic range occurring in high pH footslope
soils in southern India (Murali et al., 1978), in
waterlogged horizons in typic Cryaquents in the
Bolivian Andes (Wilke & Zech, 1987) and in
Torrifluvents formed in wadi deposits under a xeric
moisture regime in Saudi Arabia (Mashhady et al.,
1980). Only one example was found where smectite
in an Entisol was shown to form by transformation
of a mica mineral, in this case celadonite in an
altered basalt (Reid et al., 1988). This occurrence is
likely to be exceptional.
Smectite may be the dominant clay mineral in
Aridisols in the United States, and Dregne (1976)
suggested that generally, environmental conditions
are conducive to smectite formation. Smectite in
arid soils has been noted in Iraq (A1 Ravi et al.,
1969), Iran (Abtahi 1977) and Saudi Arabia (AbaHuseyn et al., 1980). Convincing evidence for a
neoformational origin for smectite in saprolites
weathering under arid conditions of the Negev
desert of southern Israel was presented by Singer
(1984). With regard to the smectitic argillic
horizons in the Argid sub-order in the USA, most
evidence suggests that these are Pleistocene relics
related to a more moist climate and to eluvial
deposition in the soil (Nettleton & Peterson, 1983).
The smectite would, therefore, be regarded as being
of inherited origin when viewed from the standpoint
of current pedogenic conditions (Buol, 1965).
Smectites originating by transformation of layer

13

silicates are unlikely to be of importance in


Aridisols.
Smectites derived from inherited materials are
common in Inceptisols, e.g. in the Fluventic
Haplaquepts of the alluvial Indo-Gangetic plain
(Islam & Lotse, 1986), in the Aquepts of the
Bangladesh Holocene floodplain (Brammer &
Brinkmann, 1977), in Typic Humaquept on glacial
marine drift in north west Washington (Pevear et
al., 1984) and in Japanese paddy soils derived from
marine alluvium (Egashira & Ohtsubo, 1983).
Neoformational smectites are also common in
Inceptisols, a good example being that described
from a Haplaquept in Nigeria by Bui & Wilding
(1988) where in the prevailing aquic environment,
the silica translocated in the ferruginous soil has
promoted the formation of an Fe-rich montmorillonite. Previous work by French scientists (Trauth et
al., 1967; Paquet 1967; Tardy et al., 1974; Pedro et
al., 1978) in this area of Africa also suggested a
widespread neoformational origin for soil smectite.
There seems to be little evidence in the literature
for smectite to originate in Inceptisols by
transformation processes.
Vertisols are dominated by smectite originating
through inheritance or neoformation. Vertisols may
have formed on alluvial plains where inherited
smectite is of detrital origin such as in Sudan
(Wilson & Mitchell, 1979), Turkey (0zkan & Ross,
1979; Gfizel & Wilson, 1981) and northern Uruguay
(Rossignol, 1983). In these instances, the smectites
are all Fe-rich with a high tetrahedral charge. Similar
smectites are found in Vertisols developed upon
basic igneous rocks, such as found in Kenya (Kantor
& Schwertmann, 1974), Israel (Singer, 1971) and
Jordan (Shadfan, 1983) where the clay mineral has a
neoformational origin. Smectites originating by
transformation of micas have not been conclusively
demonstrated to occur in Vertisols although Badroui
& Bloom (1990) and Graham & Southard (1983)
have suggested the possibility. However, it would be
surprising if a mica-to-smectite conversion is able to
occur in the poorly drained conditions characteristic
of Vertisols.
Smectites are common in Mollisols, particularly
in the United States. An early study by Glenn et al.
(1960) concluded that the smectite formed
following weathering of micas and chlorite and in
a later review, Jackson (1965) concluded that
smectites in many Mid-West soils had been
formed by K depletion of di- and tri-octahedral
micas or by the silication of polymerized sheets of

14

M. J. Wilson

hydroxy-Al. In other words, the smectites are


largely pedogenic, forming by a combination of
transformation and neoformation. In a more recent
study, Laird et al. (1988) investigated smectites in
an Argialboll-Argiaquoll toposequence in Iowa to
determine whether drainage differences influenced
the chemistry and layer charge of the clay minerals.
It was concluded that there was no drainage effect
and that the smectite was likely to be a weathering
product of inherited mica. It should be noted,
however, that a comprehensive review of Canadian
soils by Kodama (1979) reported that there had
been only minimal transformation of the layer
silicate clays in the Mollisols of the Prairie
Provinces, with slight increases in expansible
clays towards the surface in only a few instances.
In Alfisols, the smectite of the argillic layer is
surmised to have formed by different mechanisms.
Where the parent material contains illite, it has been
proposed that smectite forms by a transformation
process, with vermiculite as an intermediate phase
in some Ohio Hapludalfs (Smeck et al., 1968;
Wilding et al., 1971). Generally, however, it has
been proposed that smectite increases in the more
poorly drained Aqualfs located in low-lying areas
(Allen & Fanning, 1983), presumably being of
neoformational origin. Convincing evidence for the
synthesis of smectite in a Lithic Hapludalf derived
from serpentinite, which emphasized the role of
poor drainage in the soil, was presented by lstok &
Harward (1982).
The albic horizon in Spodosols is the site of
intense weathering and numerous studies have
demonstrated that it is dominantly smectitic and
related to the weathering of mica and sometimes
chlorite. Ross & Mortland (1966) first showed that
such smectite was beidellitic, a finding since
confirmed by others (Churchman, 1980; McDaniel
et al., 1995). The aluminous nature of these clays,
as well as the detailed study of Kodama & Brydon
(1966) which showed that the smectite mineral in
albic horizons is a three-component random
interstratification of mica, vermiculite and smectite,
represents convincing evidence that the smectite
originates through transformation of the mica. A
chlorite to smectite transformation is also possible
as Spodosols with smectitic E horizons are sometimes found on parent materials which contain
chlorite but no mica such as the tephras of the
Pacific north west USA (Ugolini et al., 1991).
Smectites are usually found as a transitory phase
in Ultisols particularly where there is an aquic

moisture regime. Thus, Kantor & Schwertmann


(1974) found that smectite increased with depth in
some Kenyan Humults where it apparently formed
as an initial weathering product of basic igneous
rocks only to decompose in the surface horizons.
Smectitic Ultisols have been more extensively
studied in the USA where they may be developed
upon smectite-rich shales, for example in the
Alabama Coastal Plains (Karathanasis et al., 1986)
and in Texas (Carson & Dixon, 1972). The above
examples illustrate that Ultisol smectites are in the
main inherited from suitable parent materials or
may form, often ephemerally, through neoformation
processes at the base of the profile. An origin
through transformation is unlikely, although this has
been suggested where Ultisols have formed on
glauconitic sediments in the Mississippi Coastal
Plain (Nash et al., 1988).
In view of the intensively weathered nature of
Oxisols, it would be expected that smectites would
be unlikely to occur. Certainly, this appears to be
true for the diagnostic oxic horizon or for the
plinthite found at depth. However, when lateritic
profiles are studied in detail, smectite is sometimes
found in the saprolite as an ephemeral phase (Nahon
& Colin, 1982; Singh & Gilkes, 1993) or even in the
so-called pallid zone (Singh & Gilkes, 1991).
Considering the relative importance of the
various modes of origin for smectite in Soil
Taxonomy, the author's conclusions are summarized in Table 3. Inheritance from smectitic parent
material is judged to be probably of greatest
importance where smectites are found in Entisols
and Ultisols and of major importance in Aridisols,
Inceptisols and Vertisols. A neoformation origin is
of major importance in Inceptisols, Alfisols,
Aridisols and Vertisols and of moderate importance
in Entisols and Mollisols. Smectite formed by
transformation is thought to be of major importance
only in Spodosols and perhaps also in Mollisols and
Alfisols although the evidence does not appear to
be entirely conclusive. As a general comment, it
seems to the author that too much importance may
have been placed upon transformation mechanisms
as a means of explaining the origin of soil smectites
and not enough on neoformation and inheritance.
Chlorites and interlayered minerals

As indicated by Barnhisel & Bertsch (1989) there


is no doubt that the occurrence in soils of primary
chlorite of ferromagnesian, trioctahedral nature, is

15

Origin and formation o f clay minerals in soils'

TABLE 3. Origin of soil smectites in soil orders as assessed from the literature.
Soil orders

Inheritance

Neoformation

Transformation

Entisols
Aridisols
Inceptisols
Vertisols
Mollisols
Alfisols
Spodosols
Ultisols
Oxisols

+++
+++
+++
+++
+
+
+++
-

++
+++
+++
+++
++
+++
+
+

+
++
+
++
++
+++
-

+++
++
+
-

Major importance
Moderate importance
Minor importance
No importance

due largely to inheritance. Primary chlorites are


generally considered to be easily weatherable and
Ross & Kodama (1976) provided convincing
experimental evidence to show that chlorite could
be changed into a regularly interstratified chloritevermiculite through chemical oxidation and selective extraction of the interlayer hydroxide sheet.
Field occurrences of the vermiculitization of
chlorite have been documented by Proust (1982),
Proust et al. (1986) and by Buurman et al. (1988).
Alternatively, chlorite may be completely decomposed without the formation of any crystalline
product, as found by Bain (1977) in podzolic soils
in Scotland.
The origin of so-called pedogenic chlorite and
interlayered or inter-gradient minerals appears to be
due entirely to transformation reactions involving
the introduction of non-exchangeable hydroxy-Al
polymers into the interlamellar space of pre-existing
smectite, vermiculite or interstratified expansible
minerals. The source of the interlayer A1 could be
direct involving acid attack of both tetrahedral and
octahedral sheets of the layer silicate mineral, or
indirect fi'om the soil solution following weathering
of aluminous minerals, like feldspars, or from
decomposition of organic matter containing
adsorbed A1. Both pathways, as well as the
conditions under which they operate, have been
confirmed by laboratory experiments. For example,
it has long been known that acidified smectite
spontaneously converts to an Al-saturated form
whereby H + is replaced by structural A1. Again, the
decomposition of trioctahedral micas by mineral
acids and non-complexing organic acids was shown

by Vicente et al. (1977) to result in the formation of


A1 hydroxy interlayered vermiculite, whereas the
effect of highly complexing organic acids often
resulted in complete breakdown of the mineral
structure. Many experiments have shown that
water-soluble polynuclear A1 complexes are
strongly adsorbed by expansible minerals to yield
hydroxy-interlayered clays (e.g. de Villiers &
Jackson, 1967), and that the original exchangeable
cations are quantitatively displaced (Brown &
Newman, 1973). According to Vicente et al.
(1977) A1 hydroxy-interlayered clays in soils
mainly relate directly to the decomposition of the
layer silicate mineral itself. They further concluded
that interlayered minerals are the most common
clay minerals in the soils of Northern Europe and
North America and that they may result not only
from natural soil processes but may also be
promoted by the application of fertilizers and the
deposition of acidity from the atmosphere.
Extensive consideration has been given to these
minerals in the latter context, particularly because
the stability of the interlayers is pH dependent,
being favoured in the pH range extending
approximately from 4.0 to 5.8 (Barnhisel &
Bertsch, 1989). Hydroxy interlayered clays can
thus serve as sources and sinks of mobile A1
depending upon soil pH conditions. Other soil
factors that promote their formation include low
organic matter content, oxidizing conditions and
frequent wetting and drying cycles. These conditions are met in a wide range of soils, and
interlayered minerals are thus commonly encountered. They may be most typical of highly

16

M. J. Wilson

weathered soils and they certainly occur in Ultisols


(Karathanasis, 1988) and even in Oxisols.
However, some emphasis has recently been
placed on the occurrence of hydroxy interlayered
clays in the podzolic soils of catchments on slowly
weathering parent rocks and where surface waters
are considered to be at risk from acidification. For
example, April et al. (1986) described hydroxy A1
interlayered vermiculite, primarily in B horizons of
soils from 26 Adirondack catchments. Olson (1988)
suggested that such minerals play a key role in
neutralizing potential increases in acid deposition in
catchments in montane areas of Maryland and
Virginia. In Scotland, Bain et al. (1990) presented
evidence to show the occurrence of these minerals
in the podzolic soils of three upland catchments and
also demonstrated the pH dependency of the
interlayers. The extent of the interlayering declined
with falling pH and complete removal was detected
below pH 4.3 in the E horizons.
Whether acid precipitation has a major role in
bringing about such changes is rather difficult to
determine, although from laboratory experiments it
is known that changes in hydroxy interlayering of
expansible phases can occur very rapidly. In most
soil studies there is no historical time scale to
assess the rate of change and to judge whether
changes coincide with marked periods of acid
deposition.
Hydroxy AI interlayers in expansible clay
minerals may become silicated as a result of
interactions either outside or within the interlamellar space (Lou & Huang, 1993) leading to the
formation of an intergradient vermiculite-kaolin
mineral (Wada & Kakuto, 1983) or to kaolinite
directly (Karathanasis & Hajek, 1983) as will be
discussed below, lnterlayered minerals are mainly
dioctahedral, representing the nature of the
precursor expansible minerals, but Al interlayered
trioctahedral vermiculites may also occur during
biotite weathering (Kato, 1965; Wilson, 1965,1966).
lnterstratified minerals

These minerals primarily represent intermediate


transformation products, mainly involving mica,
chlorite and an expansible phase, either smectite or
vermiculite (Sawhney, 1989), although there are
many examples of interstratified minerals in soils
originating by inheritance. Both regularly and
randomly interstratified clays either of dioctahedral
or trioctahedral nature, may occur in a wide variety

of soil types. However, because they represent


intermediate transformation products, they would
perhaps not be expected to occur widely in old
highly weathered soils.
Wilson & Nadeau (1985) reviewed the types of
interstratified minerals that are characteristic of
weathering environments. Hydrobiotite, consisting
of large crystals of regularly alternating biotite and
vermiculite layers and derived largely from the
weathering of biotite, represents a classic case of an
interstratified mineral that was observed early on. It
has been found many times in soils (e.g. Coleman
et al., 1963; Wilson, 1970; Kapoor, 1972), appears
to form relatively rapidly (Wilson & Nadeau,
1985), and can be synthesized in the laboratory
using dilute solutions to exchange interlayer K
(Rausel Colom et al., 1965) or by oxidation of
structural Fe during cation exchange (Farmer &
Wilson, 1970). Norrish (1973) proposed a convincing mechanism for the formation of hydrobiotite
involving sympathetic re-orientation of the hydroxyls on either side of the octahedral sheet. Removal
of K from one side of the silicate layer increases
the angle of the (OH) bond direction on that side of
the octahedral sheet, but causes a decrease in the
angle of the (OH) band on the other side. This
results in the interlayer K on this side being placed
in a more negatively charged and hence more stable
environment and alternating interlayer regions are
built up within single crystals where K is either
depleted or held very strongly. Regularly interstratified dioctahedral mica-vermiculite is also
found in soil clay fractions particularly in the
upper horizons of podzolic soils (Churchman,
1978,1980; Wilson et al., 1984) and presumably
formed from dioctahedral mica by a mechanism
similar to that of hydrobiotite formation. However,
the dioctahedral mineral does not seem to occur in
non-clay fractions. Further weathering of vermiculitized mica results in a lower negative charge on
the silicate sheet leading to the formation of
smectitic products of both a trioctahedral (Ismail,
1969; MacEwan, 1954) and dioctahedral nature
(Churchman, 1980). In the latter instance a
regularly interstratified mineral may form with the
XRD characteristics of rectorite.
Regularly interstratified chloritic material may
occur in soils as a consequence of the vermiculitization of chlorite, although the mechanism involved is
not clear. A regularly interstratified chloritevermiculite yielding a high spacing of 28 A with
many lower orders was described by Johnson

Origin and formation o f clay minerals' in soils

(1964) from the C horizon of a soil developed on


metamorphosed basalt in Pennsylvania and a similar
mineral was reported by Herbillon & Makumbi
(1975) in a recent tropical soil derived from
chlorite-schist in Za'ire. In the latter instance, at
least, it was evident that the chlorite from which the
interstratified product developed was already partly
vermiculitized (Makumbi & Herbillon, 1972) but
the experiments of Ross & Kodama (1976)
demonstrate that some true chlorites will break
down to a regularly interstratified product.
The formation of regularly interstratified minerals
and thence of vermiculite or smectite must involve
transition through randomly interstratified phases
but it is of interest that these phases appear to be
relatively transitory and do not seem to persist in
soil profiles. Comparison with diagenetic sequences
is striking where great thicknesses of argillaceous
sediment may be characterized by randomly
interstratified illite-smectite which converts slowly
with depth and increasing temperature and pressure
to more ordered forms including rectorite-like
structures (R1) and more long-range ordered
structures (R2 and R3). In the author's view, there
is little evidence that the long-range ordered illitesmectites nor indeed the precursor randomly
interstratified illite-smectite clay form as stable
phases in soils. This is perhaps not surprising in
view of the increasing P - T conditions required for
this transformation in diagenesis. One of the few
papers to show that interstratified illite-smectite will
form at earth surface conditions is that of Bergkraut
et al. (1994) in a study of weathering in basic
pyroclastics. However, the illite-smectite differs
from that found in diagenetic sequences in that it
is Fe-rich, recalling the illitic clays formed at earth
surface conditions.
Interstratified kaolin-smectite is now becoming
more widely reported in soils. It appears to be
particularly common in soils derived from basic
igneous rocks and, for example, was identified by
Norrish & Pickering (1983) as the only clay mineral
in ~40 Australian soils developed upon such parent
material. The occurrence of kaolinite-smectite in a
red-black toposequence in Burundi (Herbillon et al.,
1981) suggests a genetic link between the black
smectitic soils in the poorly drained footslopes and
the red kaolinitic soils on the freely-drained hill
tops. More recent work shows that the kaolin
mineral in this kind of interstratification may be
halloysite (Delvaux et al., 1990) and that the
expansible mineral can be vermiculitic (Btihmann

17

& Grubb, 1991). The transformation of smectite to


interstratified halloysite-smectite was shown in
weathered acid clay in Japan to involve the
crystallization of excess silica in a separate phase
(Watanabe et al., 1992).
Although Wilson (1987) concluded that the
concept of fundamental particles was probably not
of general applicability in the interpretation of
interstratified smectitic clays in soils, Robert et al.
(1991) showed by HRTEM that these particles did
exist in soils and, from these observations, Aouidjit
et al. (1996) distinguished two types of interstratification described as 'structural' and 'textural'.
The former involves relatively large crystals of
micrometric lateral extent, variable thickness and
three-dimensional structure. Textural interstratification consists of associations of elementary particles
with a random two-dimensional structure and is of
variable thickness and lateral extent.
Kaolin minerals

Dixon (1989) reviewed the origin and formation


of kaolin minerals and showed that precipitation
from solution of kaolinite required acid conditions
with moderate silica activity and small amounts of
base cations. Kaolinite was synthesized fi-om
hydroxy A1 interlayered montmorillonite by
Poncelet & Brindley (1967) but only under
hydrothermal conditions. Neoformation and transformation mechanisms are, therefore, feasible.
Kaolinite is generally found as a minor constituent
in young soils, such as those derived from glacial
material, and by implication is considered to require
>10,000 years to form. It may be an abundant
constituent of soils developed upon old geomorphic
surfaces.
Compared with the kaolinites of geological
deposits, soil kaolinites are often of a smaller
particle size, tend to be highly disordered and may
contain Fe as an isomorphous substituent. They
may also be interstratified with smectites. The
pioneering work of Angel et al. (1974), Jones el al.
(1974) and Meads & Malden (1975) using EPR
spectroscopy showed that Fe3+ could be substituted
for A1 in the octahedral sheet of kaolinite and that
there were different sites of substitution. For the
kaolinite of tropical soils Herbillon et al. (1976)
demonstrated a relationship between Fe content,
crystallinity index and EPR spectral features,
particularly the intensity and asymmetry of the
gef,~4 band. Mendelovici et al. (1979) and

18

M. J. Wilson

Cantinolle et al. (1984) showed convincingly by IR


spectroscopy that structural Fe occurs in kaolinite in
lateritic and bauxitic soils. It seems, therefore, that
soil kaolinites (like soil smeetites), may be different
from the kaolinites of geological deposits, presumably because of the large amounts of Fe in the
pedogenic environment combined with conditions
that promote the mobilization of this element, at
least on a local scale. The characterization of
kaolinite by EPR was developed further by Miiller
& Calas (1993) who showed that the technique
could be used to distinguish different environments
of formation according to total Fe 3+ content,
distribution among different sites (Fe(~) and
Fe(ii)), concentration and types of paramagnetic
defect centres (PDC) and presence or absence of
sorbed species. Soil kaolinites yielded a high
intensity Fe 3+ EPR signal, showed an interdependence between Fe(~) concentrations and crystalline
order, low intensity signals from PDCs and lack of
resonance due to sorbed species.
Halloysite occurs primarily in youthful volcanicderived soils, but it also forms from primary
minerals in tropical soils or pre-glacially weathered
materials. Most commonly the mineral is of a
tubular or spheroidal morphology. Platy halloysite
is rare. Although similar to kaolinite in its layer
structure, it is important to realize that halloysite is
fundamentally different in that it has a 2-layer
structure whereas kaolinite has a l-layer structure
(Bailey, 1989). This means that a halloysite to
kaolinite conversion cannot take place by a simple
structural transformation but would require a
recrystallization process, as was found experimentally by La Iglesia & Gal~in (1975). The reasons for
the formation of spheroidal halloysite as opposed to
tubular halloysite are not clear, although it may be
that high levels of silica supersaturation are
required (Tomura et al., 1985) and the appropriate
physical conditions provided by pumice grains.
Tubular halloysite may eventually form from the
spheroidal form (Sudo & Yotsumoto, 1977). It is
worth noting too that, like kaolinite, halloysite may
contain structural Fe (Wada & Mizota, 1982),
particularly where it has formed from the weathering of ferruginous material (Dao Cho & Mermut,
1992). There is evidence gained by X-ray photoelectron spectroscopy that spheroidal halloysite may
have a greater Fe content than the tubular form
(Bailey, 1989).
Although most studies would seem to indicate
that kaolinite in soils has formed by neoformation,

there is evidence that a transformation mechanism


is feasible. Thus, Wada & Kakuto (1983) presented
evidence to show that in Korean Ultisols, kaolinite
had formed via an intergradient vermiculite-kaolin
mineral involving tetrahedral inversion and attachment to the hydroxy A1 interlayer sheet. A similar
mechanism for the formation of kaolinite from
hydroxy interlayered smectite was proposed by
Karathanasis & Hajek (1983).
FUTURE

PERSPECTIVES

This overview has highlighted the necessity for


further research to address many specific questions
and gaps in our knowledge as to the origin and
formation of clay minerals in soils, as well as
further characterization of the detailed nature of the
minerals themselves.
With regard to soil micas, it is clear that further
detailed studies are needed on their crystallochemistry and on the changes that occur with decreasing
particle size, particularly the possible conversion
from the 2M1 to the 1Md polytype and the role of
microdivision in bringing about fundamental
particles, or at least particles that are more
susceptible to vermiculitization. At this stage the
relative roles of increased basal surface as opposed
to edge surface is unclear. The question of
formation of pedogenic micas is one that remains
to be resolved. Even the reality of the process
seems uncertain, a point that could be clarified by
K/Ar dating and other studies of carefully
fractionated samples.
For reasons that have been given above, most of
the information that we have on vermiculite refers
to trioctahedral vermiculite, despite the fact that
dioctahedral vermiculite is much more common in
soils. Therefore, we need to know more about the
mechanisms of formation of dioctahedral vermiculites, particularly the changes in chemistry involved
in the reduction of layer charge during evolution
from the parent micas. In this instance, oxidation of
structural Fe is unlikely to play as significant a role
as it does in trioctahedral vermiculites. There is
very little information on the stacking sequence of
the silicate layers and the organization of the
interlamellar space for dioctahedral vermiculites,
in comparison with that available for their
trioctahedral counterparts.
Although there is good evidence that soil
smectites may form through both transformation
and neoformation pathways, more information is

Origin and formation o f clay minerals in soils

needed with regard to possible differences in


chemistry and structure that these two modes of
origin may engender. For example, transformation
smectites should, in principle, be beidellitic, as they
should inherit Al-substituted tetrahedral sheets from
their parent micas, and they might also be expected
to show semi-random stacking (again an inherited
feature from their precursor vermiculites), as
opposed to turbostratic stacking. The mechanism
involved in the reduction of layer charge, and the
distinction from the vermiculite from which the
transformation smectites are derived is another area
for further research. With regard to neoformation
smectites, confirmation is required by direct
methods that many can indeed be described as
ferruginous beidellites, as is suggested by chemical
analyses. However, such analyses can be misleading
because non-structural Fe oxides may be resistant to
deferration techniques. There may perhaps be scope
to use EPR techniques for the assessment of
environmental conditions for smectite formation,
in the same way as they have been used in kaolinite
studies. In any event, it would seem that detailed
comparison between transformation and neoformation soil smectites would be worthwhile.
In the author's view, there is ample scope for
further research into the interstratified clay minerals
that are truly characteristic of the pedogenic
environment. It seems that the present picture is
somewhat confused by the implied formation in
soils of the same types of interstratified illitesmectites that are so common in the diagenetic
environment. This is unreasonable, bearing in mind
that the latter are characterized by increasing
conditions of temperature and pressure and that
there is a dearth of evidence to show that
diagenetic-type illite-smectites really do form in
the soil. It is true that K saturation combined with
wetting and drying cycles can convert smectite to
an illite-like product in the laboratory, but this
product is not illite sensu stricto and K saturation is
unlikely to occur in the soil except under arid
conditions. Further research is required to address
this issue. Other specific problems that might be
addressed include the mechanisms for the formation
of regularly interstratified dioctahedral mica-vermiculite and for interstratified kaolinite-smectite and
halloysite-smectite in soils. Further information is
required too on the possible role of microdivision
and the creation of fundamental particles in
promoting interstratification as suggested by
Robert and his colleagues.

19

For the kaolin minerals, it seems that genesis


through a transformation mechanism from hydroxyinterlayered minerals involving tetrahedral inversion
of silicate layers requires further documentation and
assessment as to its general applicability in soils.
Also, there is no question that a better understanding is required of the conditions of formation
of halloysite as opposed to kaolinite, in soils, and in
particular the formation of spheroidal halloysite vs.
tubular halloysite. In these endeavours, it seems
certain that the use of advanced spectroscopic
techniques such as EPR will play an ever increasing
role.
With regard to the global overviews of clay
mineral formation in soils described at the
beginning of this paper, it still seems that they
can be regarded as generally valid and that they
provide a useful frame of reference. It may be,
however, that uncertainty still remains about the
relative importance of transformation and neoformation in both general and specific instances.
Figure 2 summarizes the author's views on this
matter in terms of primary and secondary pathways
for clay mineral formation in soils in a global
context. It can be seen that soil smectites and
kaolinites are regarded primarily as the products of
neoformation. For smectite, an important, but
s e c o n d a r y p a t h w a y , a p p l i c a b l e m a i n l y to
Spodosols and perhaps to Mollisols and Alfisols,
is through the weathering of inherited mica (and to
a lesser extent chlorite) and may or may not involve
the intervention of hydroxy-A1 interlayered
minerals. Kaolinite too, although primarily a soil
mineral arising from neoformation, also forms
through other pathways of transformation, namely
direct weathering of mica, conversion of hydroxyA1 interlayered vermiculite or smectite and weathering of smectite through an intermediate kaolinitesmectite interstratified phase. Halloysite is regarded
primarily as a product of neoformation, although it
seems that it may also form through an interstratified halloysite-smectite as a result of smectite
weathering.
Finally, it is worth emphasizing that the question
of the origin and formation of clay minerals in soils
is not an exercise of mere academic interest but
may relate directly to soil properties and behaviour
in numerous ways. A better understanding of the
detailed nature of clay minerals in soil and the way
in which they relate to overall soil chemical and
physical properties, combined with a knowledge of
their mode of formation, provides a sound basis for

20

M. J. Wilson

TRANSFORMATION
MICA
(chlorite)

INTERSTRATIFIED

'

VERMICULITE.
ALUMINIUM

SMECTITE

'

INTERLAYERED' ~ ~ ' ' ' ~ i

INTERSTRATIFIED
KAOLIN-SMECTITE

~ KAOLINITE

NEOFORMATION
FIG. 2. Pathways for the formation of clay minerals in soils. Note that the major pathways for smectite and
kaolinite are considered to be through neoformation as represented by thick arrows. Transformation pathways to
smectite occur via interstratified and interlayered phases. A direct conversion of mica to kaolinite can also occur
but this may not be regarded as a true transformation because of the structural differences between the minerals,
hence the dotted line.

being able to predict, at least in general terms, soil


behaviour in the context of both agricultural and
environmental problems.

ACKNOWLEDGMENTS
I am indebted to the Clay Minerals Group Committee
for the invitation to prepare this review paper. This
work was supported by the Scottish Office Agriculture
Environment and Fisheries Department.

REFERENCES
Aba-Huseyn M.M., Dixon J.g. & Lee S.Y. (1980)
Mineralogy of Saudi Arabian soils: south western
Region. Soil Sci. Soc. Am. J. 44, 643-649.
Abtahi A. (1977) Effect of saline and alkaline ground
water on soil genesis in semi arid Southern Iran. Soil
Sci. Soc. Am. Jr. 41, 583 588.
Allen B.L. & Fanning D.S. (1983) Comparison and Soil
Genesis. Pp. 141-192 in: Pedogenesis and Soil
Taxonomy. L Concepts" and Interactions (L.P.

Origin and formation of clay minerals' in soils


Wilding, N.E. Smeck & G.F. Hall, editors). Elsevier,
Amsterdam.
A1 Ravi A.H., Jackson M.L. & Hole F.D. (1969)
Mineralogy of some arid and semi-arid land soils
of Iraq. Soil Sci. 107, 480-486.
Allen B.L. & Hajek B.F. (1989) Mineral Occurrence in
Soil Environments. Pp. 199-278 in: Minerals in Soil
Environments' (J.B. Dixon & S.B. Weed, editors).
Soil Sci. Soc. America, Madison, Wisconsin, USA.
Angel B.R., Jones J.P.E. & Hall P.L. (1974) Electron
spin resonance studies of doped synthetic kaolinite.
I. Clay Miner. 10, 247-255.
Aouidjit H., Elsass F., Righi D. & Robert M. (1996)
Mica weathering in acidic soils by analytical electron
microscopy. Clay Miner. 31, 319 332.
April R.H., Hluchy M.M. & Newton R.M. (1986) The
nature of vermiculite in Adirondack soils and till.
Clays Clay" Miner. 34, 549-556.
Badraoui M. & Bloom P.R. (1990) Iron rich high charge
beidellite in Vertisols and Mollisols of the High
Chaouia Region of Morocco. Soil Sci. Soc. Am. J. 54,
267 -274.
Bailey S.W. (1989) Halloysite - - a critical assessment.
Proc. 9th Int. Clay Conf Strasbourg, 89-98.
Bain D.C. (1977) The weathering of ferruginous chlorite
in a podzol from Argyllshire, Scotland. Geoderma,
17, 193-208.
Bain D.C., Mellor A. & Wilson M.J. (1990) Nature and
origin of an aluminous vermiculitic weathering
product in acid soils from upland catchments in
Scotland. Clay Miner. 25, 467-475.
Barnhisel R.I. & Bertsch P.M. (1989) Chlorites and
Hydroxy interlayered Vermiculite and Smectite. Pp.
729-788 in: Minerals" in Soil Environments (J.B.
Dixon & S.B. Weed, editors). Soil Sci. Soc. Amer.,
Madison, Wisconsin.
Barshad i. & Kishk F.M. (1969) Chemical composition
of soil vermiculite clays as related to their genesis.
Contrib. Mineral. Pet. 24, 136-155.
Basham I.R. (1974) Mineralogical changes associated
with deep weathering of gabbro in Aberdeenshire.
Clay Miner. 10, 189 202.
Bergkraut V., Singer A. & Stahr K. (1994) Palagonite
reconsidered: paracrystalline illite-smectite from
regoliths on basic pyroclastics. Clays Clay Miner.
42, 582-592.
Borchardt G. (1989) Smectites. Pp. 675-727 in:
Minerals in Soil Environments (J.B. Dixon & S.B.
Weed, editors). Soil Sci. Soc. America Madison,
Wisconsin.
Brammer H. & Brinkman R. (1977) Surface water gley
soils in Bangladesh: Environment landforms and soil
morphology. Geoderma, 17, 91-109.
Brown G. (1953) The dioctahedral analogue of vermiculite. Clay Miner. Bull. 2, 64-69.
Brown G. & Newman A.C.D. (1973) The reactions of
soluble aluminium with montmorillonite. J. Soil Sci.

21

24, 339-354.
Bfihmann C. & Grubb P.L.C. (1991) A kaolinite
smectite interstratification sequence from a red-black
toposequence. Clay Miner. 26, 343-358.
Bui E. & Wilding L.P. (1988) Pedogenesis and
mineralogy of a Haplaquept in Niger (West
Africa). Geoderma, 43, 49-64.
Buol S.W. (1965) Present soil-forming factors and
processes in arid and semi arid regions. Soil Sci.
99, 45-49.
Buurman P., Meyer E.L. & Van Wijk J.H. (1988)
Weathering of chlorite and vermiculite in ultramafic
rocks of Cabo Ortegal, northwest Spain. Clays Clay
Miner. 36, 263-269.
Calle de la C. & Suquet H. (1988) Vermiculite. Pp.
455-496 in: Hydrous Phyllosilicates (exclusive of
micas') (S.W. Bailey, editor). Vol. 19. Reviews in
Mineralogy. M i n e r a l o g i c a l Soc. A m e r i c a ,
Washington.
Cantinolle P., Didier P., Meunier J.D., Parron C.,
Guendon J.L., Bocquier G. & Nahon D. (1984)
Kaolinites ferriferes et oxy-hydroxydes de fer et
d'alumine dans les bauxites des Cantonnettes (S E de
la France). Clay Miner. 19, 125-135.
Carson C.D. & Dixon J.B. (1972) Potassium selectivity
in certain montmorillonitic soil clays. Soil Sci. Soc.
Am. Proc. 36, 838-843.
Churchman G.J. (1978) Studies on a climax sequence in
soils in tussock grasslands. Mineralogy N. Z. J. Sci.
21, 467-480.
Churchman G.J. (1980) Clay minerals formed from
micas and chlorites in some New Zealand soils. Clay
Miner. 15, 59 76.
Coleman N.T., LeRoux F.H. & Cady J.G. (1963)
Biotite-hydrobiotite-vermiculite in soils. Nature
Lond. 198, 409-410.
Dao Cho Lt. & Mermut A.R. (1992) Evidence for
halloysite formation from weathering of ferruginous
chlorite. Clays' Clay Miner. 40, 608-619.
De Kimpe C. & Tardy Y. (1968) Etude de l'alteration
d'une biotite en kaolinite par spectroscopic infrarouge. Bull. Groupe Franc. Argiles, 19, 81-85.
Delvaux B., Herbillon A.J., Vielvoye L. & Mestdagh
M.M. (1990) Surface properties and clay mineralogy
of hydrated halloysite soil clays. II. Evidence for the
presence of halloysite smectite (H/Sm) mixed layer
clays. Clay Miner. 25, 141-160.
Dixon J.B. (1989) Kaolin and Serpentine Group
Minerals. Pp. 467 525 in: Minerals in Soil
Environments (J.B. Dixon & S.B. Weed, editors).
Soil Sci. Soc. America, Madison, Wisconsin.
Dixon J.B. & Weed S.B. (1989) Minerals" in Soil
Environments'. Soil Sci. Soc. America, Madison,
Wisconsin.
Douglas L.A. (1989) Vermiculites. Pp. 635-674 in:
Minerals' in Soil Environments' (J.B. Dixon & S.B.
Weed, editors). Soil Sci. Soc. America, Madison,

22

M. J. Wilson

Wisconsin.
Dregne H.E. (1976) Soils' o f Arid Regions. Elsevier,
New York.
Duchaufour P. (1960) Prdcis de Pedologie. Masson,
Paris.
Dymond J., Biscaye P.E. & Rex R.W. (1974) Eolian
origin of mica in Hawaiian soils. Geol. Soc. Am.
Bull. 85, 37-40.
Egashira K. & Ohtsubo M. (1983) Swelling and
mineralogy of smectites in paddy soils derived from
marine alluvium, Japan. Geoderma, 29, 119-127.
Fanning D.S., Keramidas V.Z. & E1-Desoky M.A.
(1989) Micas. Pp. 551-634 in: Minerals in Soil
Environments (J.B. Dixon & S.B. Weed, editors).
Soil Sci. Soc. America, Madison, Wisconsin.
Farmer V.C. & Wilson M.J. (1970) Experimental
conversion of biotite to hydrobiotite. Nature Lond.
226, 841-842.
Farmer V.C., Russell J.D., McHardy W.J., Newman
A.C.D., Alrichs J.L. & Rimsaite J.Y.H. (1971)
Evidence for loss of protons and octahedral iron
from oxidized biotites and vermiculites. Mineral.
Mag. 38, 121 137.
Fordham A.W. (1990a) Formation of trioctahedral illite
from biotite in a soil profile over granite gneiss.
Clays' Clay Miner. 38, 187-195.
Fordham A.W. (1990b) Weathering of biotite into
dioctahedral clay minerals. Clay Miner. 25, 51 63.
Glenn R.C., Jackson M.L., Hole F.D. & Lee G.B. (1960)
Chemical weathering of layer silicate clays in loessderived Tama silt loam of southwestern Wisconsin.
Clays Clay Miner. 8, 63-83.
Graham R.C. & Southard A.R. (1983) Genesis of a
Vertisol and an associated Mollisol in Northern
Utah. Soil Sci. Soc. Am. J. 47, 552 559.
Giizel N. & Wilson M.J. (1981) Clay mineral studies of
a soil chronosequence in southern Turkey.
Geoderma, 25, 113 129.
Herbillon A.J. & Makumbi M.V. (1975) Weathering of
chlorite in a soil derived from a chlorite-schist under
humid tropical conditions. Geoderma, 13, 89-104.
Herbillon A.J., Mestdagh M.M., Vielvoye L. &
Derouane E.G. (1976) Iron in kaolinites with special
reference to kaolinites from tropical soils. Clay
Miner. 11,201-220.
Herbillon A.J., Frankort R. & Vielvoye L. (1981) An
occurrence of interstratified kaolinite-smectite
minerals in a red-black soil toposequence. Clay
Miner. 16, 195 201.
Islam A.K.M.E. & Lotse E.G. (1986) Quantitative
mineralogical analysis of some Bangladesh soils
with X-ray, ion exchange and selective dissolution
techniques. Clay Miner. 21, 31-42.
Ismai/G.T. (1969) Role of ferrous iron oxidation in the
alteration of biotite and its effect on the type of clay
minerals formed in soils of arid and humid regions.
Am. Miner. 54, 1460 1466.

Istok J.D. & Harward M.E. (1982) Influence of soil


moisture on smectite formation in soils derived from
serpentinite. Soil Sci. Soc. Am. J. 46, 1106 1108.
Jackson M i . (1964) Chemical composition of soils. Pp.
71-141 in: Chemistry o f the Soil (F.E. Bear, editor).
Reinhold Publishing Corp., New York.
Jackson M.L. (1965) Clay transformations in soil
genesis during the Quaternary. Soil Sci. 99, 15 21.
Johnson J.J. (1964) Occurrence of regularly interstratifled chlorite-vermiculite as a weathering product of
chlorite in a soil. Am. Miner. 49, 556-572.
Jones J.P.E., Angel B.R. & Hall P.L. (1974) Electron
spin resonance studies of doped synthetic kaolinite.
II. Clay Miner. 10, 257-270.
Juang T.C. & Uehara G. (1968) Mica genesis in
Hawaiian soils. Soil Sci. Soc. Am. Proc. 32, 31-35.
Juo J.S.R. & White J.L. (1969) Orientation of the dipole
moments of hydroxyl groups in oxidized and
unoxidized biotite. Science, 165, 804-805.
Kantor W. & Schwertmann U. (1974) Mineralogy and
genesis of clays in red-black soil toposequences on
basic igneous rocks in Kenya. J. Soil Sci. 25, 63 78.
Kapoor B.S. (1972) Weathering of micaceous clays in
some Norwegian podzols. Clay Miner. 9, 383 394.
Karathanasis A.D. (1988) Compositional and solubility
relationships between aluminium-hydroxy interlayered soil smectites and vermiculites. Soil Sci.
Soc. Am. J. 52, 1500 1508.
Karathanasis A.D. & Hajek B.F. (1984) Evaluation of
aluminium-smectite equilibria in naturally acid soils.
Soil Sci. Soc. Am. J. 48, 413 417.
Karathanasis A.D., Hurt G.W. & Hajek B.F. (1986)
Properties and classification of montmorillonite-rich
Hapludults in the Alabama coastal plains. Soil Sci.
142, 76 82.
Kato K. (1965) Mineralogical study of weathering
products of granodiorite at Shinshiro city (V)
Trioctahedral aluminium vermiculite as a weathering
product of biotite. Soil Sci. Plant Nutrition, Tokyo,
11, 114 122.
Kodama H. (1979) Clay minerals in Canadian soils:
their origin, distribution and alteration. Can. J. Soil
Sci. 59, 37 58.
Kodama H. & Brydon J.E. (1966) lnterstratified
montmorillonite-mica clays from sub-soils of the
Prairie Provinces, Western Canada. Clays Clay
Miner. 13, 151-173.
La Iglesia A. & Galfin E. (1975) Halloysite-kaolinite
transformation at room temperature. Proc. Int. Clay
Conf. Madrid, 173-185.
La Manna J.M. & Ugolini F.C. (1987) Trioctahedral
vermiculite in a 1980 pyroclastic flow, Mt St Helens,
Washington. Soil Sci. 143, 162-167.
Laird D.A. & Nater E.A. (1993) Nature of the illitic
phase associated with randomly interstratified smectite-illite in soils. Clays Clay Miner. 41,280-287.
Laird D.A., Fenton T.E. & Scott A.D. (1988) Low

Origin and formation o f clay minerals in soils


charge of smectites in an Argialboll-Agriaquoll
Sequence. Soil Sci. Soc. Am. J. 52, 4630467.
Lou G. & Huang P.M. (1993) Silication of hydroxy
aluminium interlayers in smectite. Clays Clay Miner.
41, 38 44.
MacEwan D.M.C. (1954) 'Cardenite', a trioctahedral
montmorillonoid derived from biotite. Clay Miner.
Bull. 2, 120-126.
Mahjoory R.A. (1975) Clay mineralogy, physical and
chemical properties of some soils in arid regions of
Iran. Soil Sci. Soc. Am. Proc. 39, 1157-1164.
M a k u m b i M . N . & H e r b i l l o n A.J. ( 1 9 7 2 )
Vermiculitisation 6xperimentale d'une chlorite. Bull
Groupe Franc. Argiles, 24, 153-164.
Martin-Garcia J.M., Delgado G., Sfindez-Maro~rn M.,
Pfirraga J.F. & Delgado R. (1997) Nature of
dioctahedral micas in Spanish red soils. Clay
Miner. 32, 107-122.
Masshady A.S., Reda M., Wilson M.J. & Mackenzie
R.C. (1980) Clay and silt mineralogy of some soils
from Qasim, Saudi Arabia. J. Soil Sei. 31, 101 115.
McDaniel P.A., Falen A.L., Tice K.R., Graham R.C. &
Fendorf S.E. (1995) Beidellite in E horizons of
Northern Idaho spodosols formed in volcanic ash.
Clays' Clay Miner. 43, 525 532.
Meads R.E. & Malden P.J. (1975) Electron spin
resonance in natural kaolinites containing Fe 3+ and
other transition metal ions. Clay Miner. 10,
313 345.
Mendelovici E., Yariv Sh. & Villalba R. (1979) Ironbearing kaolinite in Venezuelan laterites. I. Infrared
spectroscopy and chemical dissolution evidence.
Clay Miner. 14, 323-331.
Millot G. (1964) Geologie des Argiles. Masson, Paris.
Muller J.P. & Calas G. (1993) Genetic conditions of
paramagnetic centers in kaolinites. Pp. 261-289 in:
Kaolin Genesis' and Utilization (H. Murray, W.
Bundy & C. Harvey, editors). Clay Miner. Soc.,
Boulder, Colorado.
Murali V., Krishna Murti G.S.R. & Sarma V.A.K.
(1978) Clay mineral distribution in two toposequences of tropical soils of India. Geoderma, 20,
257-277.
Nadeau P.H., Wilson M.J., McHardy W.J. & Tait J.
(1984) Interstratified clays as fundamental particles.
Science, 225, 923-925.
Nahon D.B. & Colin F. (1982) Chemical weathering of
orthopyroxenes under lateritic conditions. Am. J. Sci.
282, 1232-1243.
Nash V.E., Pettry D.E. & Mohd Noor Sudin (1988)
Mineralogy and chemical properties of two Ultisols
formed in glauconitic sediments. Soil Sci. 145,
270-277.
Nettleton W.D., Nelson R.E. & Flach K.W. (1973)
Formation of mica in surface horizons of dryland
soils. Soil Sci. Soc. Am. Proc. 37, 473-478.
Nettleton W.D. & Peterson F.F. (1983) Aridisols. Pp.

23

165--215 in: Pedogenesis and Soil Taxonomy. II.


The Soil Orders (L.P. Wilding, N.E. Smeck & G.F.
Hall, editors). Elsevier, Amsterdam.
Newman A.C.D. (1969) Cation exchange properties of
micas. I. The relation between mica composition and
K-exchange in solutions of different pH. J. Soil Sci.
20, 357 373.
N i e d e r b u d d e E.A. (1975) V e r a n d e r u n g e n von
Dreischicht-Tonmineralen durch natives K in holoz/inen Mittel Deutschlands und Niederbayems. Z
Pflanzennernaehr. Bodenk. 138, 217-234.
Niederbudde E.A. & Kussmaul H. (1978) Tonmineral
eigenschaften und Umwandlungen in ParabraunerdeP r o f i l p a e r e n u n t e r A c k e r u n d W a l d in
Siiddeutschland. Geoderma, 20, 239-255.
Norrish K. (1973) Factors in the weathering of mica to
vermiculite. Proc. Int. Clay C o n f , Madrid,
417-432.
Norrish K. & Picketing J.G. (1983) Clay Minerals. Pp.
2 8 2 - 3 0 8 in: Soils': an Australian Viewpoint:
Division o f Soils, CSIRO. CSIRO, Melbourne/
Academic Press, London.
Ojanuga A.G. (1973) Weathering of biotite in soils of a
humid tropical climte. Soil Sei. Soc. Am. Proc. 37,
644 646.
Olson C.G. (1988) Clay mineral contribution to the
weathering mechanisms in two contrasting watersheds. J. Soil Sci. 39, 457-468.
()zkan A. & Ross G.J. (1979) Ferruginous beidellites in
Turkish soils. Soil Sci. Soc. Am. J. 43, 1242-1248.
Paquet H. (1967) Les montmorillonites des vertisols:
alteration alcaline en milieu tropicale. Bull Serv.
Carte-Geol Als. Lorr. 20, 293-306.
Pedro G. (1964) Contribution a 1'~tude exp~rimentale de
l'alter~ration geochimique des' roches crystallines'.
Thbse Paris, Ann Agron.
Pedro G. (1982) The conditions of formation of
secondary constituents. Pp. 63-81 in: Constituents'
and Properties' o f Soils (M. Bonneau & B. Souchier,
editors) Academic Press, London.
Pedro G., Carmouze J.-P. & Velde B. (1978) Peloidal
nontronite formation in recent sediments of Lake
Chad. Chem. Geol. 23, 139-149.
Pevear D.R., Dethier D.P. & Frank D. (1982) Clay
minerals in the 1980 deposits from Mount St Helens.
Clays' Clay Miner. 30, 241-252.
Pevear D.R., Goldin A. & Sprague J.W. (1984) Mineral
transformations in soils formed in glacial marine
drift, Northwestern Washington. Soil Sci. Soc. Am. J.
48, 208-216.
Poncelet G.M. & Brindley G.W. (1967) Experimental
formation of kaolinite from montmorillonite at low
temperatures. Am. Miner. 52, 1161 - 1173.
Proust D. (1982) Supergene alteration of metamorphic
chlorite in an amphibolite from the Massif Central,
France. Clay Miner. 17, 159-173.
Proust D., Eyinery J.P. & Beaufort D. (1986) Supergene

24

M J. Wilson

vermiculitization of a magnesium chlorite: iron and


magnesium removal processes. Clays Clay Miner.
34, 572-580.
Rausel Colom J.A., Sweatman T.R., Wells C.B. &
Norrish K. (1965) Studies in the artificial weathering
of mica. Pp. 4 0 - 7 2 in: Experimental Pedology (E.G.
Hallsworth & D.V. Crawford, editors). Butterworths,
London.
Reid D.A., Graham R.C., Edinger S.B., Baurent B.H. &
Ervin J.D. (1988) Celadonite and its transformation
to smectite in an Entisol at Red Rock Canyon, Kern
County, California. Clays Clay Miner. 36, 425-431.
R o b e r t M., H a r d y M. & E l s a s s F. ( 1 9 9 1 )
Crystallochemistry, properties and organization of
soil clays derived from major sedimentary rocks in
France. Clay Miner. 26, 409-420.
Romero R., Robert M., Elsass F. & Garcia C. (1992)
Evidence by transmission electron microscopy of
weathering microsystems in soils developed from
crystalline rocks. Clay Miner. 27, 21-34.
Ross G.J. & Mortland M.M. (1966) A soil beidellite.
Soil Sci. Soc. Am. Proc. 30, 337-343.
Ross G.J. & Kodama H. (1976) Experimental alteration
of chlorite into a regularly interstratified chloritevermiculite by chemical oxidation. Clays Clay
Miner. 24, 183 190.
Rossignol J.P. (1983) Les Vertisols du nord de
l'Uruguay. Cah. ORSTOM s& Pddol. 20, 271-291.
Rutherford G.K. & Debenham P.L. (1981) The mineralogy of some silt and clay fractions from some soils
on the Faeroe Islands. Soil Sci. 132, 288-299.
Sawhney B.L. (1989) Interstratification in Layer
Silicates. Pp. 789 828 in: Minerals in Soil
Environments' (J.B. Dixon & S.B. Weed, editors).
Soil Sci. Soc. America, Madison, Wisconsin.
Shadfan H. (1983) Clay minerals and potassium status
of some soils of Jordan. Geoderma, 31, 41-56.
Singer A. (1971) Clay minerals in the soils of the
southern Golan Heights. lsr. J. Earth Sei. 20,
105 115.
Singer A. (1984) Clay formation in saprolites of igneous
rocks under semi arid to arid conditions, Negev,
Southern Israel. Soil Sci. 137, 332-340.
Singh B. & Gilkes R.J. (1991) A K-rich beidellite from a
lateritic pallid zone in West Australia. Clay Miner.
26, 233 244.
Singh B. & Gilkes R.J. (1993) Weathering of spodumene to smectite in a lateritic environment. Clays
Clay Miner. 41, 624-630.
Smeck N.E., Wilding L.P. & Hollawaychuk N. (1968)
Genesis of argillic horizons in Celina and Morley
soils of western Ohio. Soil Sci. Soc. Am. Proc. 32,
550 556.
Soil Survey Staff (1960) Soil Classification. A
Comprehensive System. USDA.
Stephen I. (1952) A study of rock weathering with
reference to the soils of the Malvern Hills. II.

Weathering of appinite and Ivy Scar rock. J. Soil Sci.


3, 219 237.
Sudo T. & Yotsumoto H. (1977) The formation of
halloysite tubes from spherulitic halloysite. Clays
Clay Miner. 25, 155-159.
Tardy Y., Cheverry C. & Fritz B. (1974) Neoformation
d'une argile magnesienne darts les d6pressions
interdunaires du lac Tchad. Applications aux domaines de stabilit6 des phyllosilicates alumineux,
magnesiens et ferriferes. C. R. Acad. Sci. Ser. D.
278, 1999 2002.
Tomura S., Shibasaki Y. & Mizuta H. (1985) Growth,
conditions and genesis of spherical and platy
halloysite. Clays Clay Miner. 33, 200 206.
Trauth N., Paquet H., Lucas J. & Millot G. (1967)
Montmorillonite of lithomorphic Vertisols are ferriferous: geochemical and sedimentological consequences. C. R. Acad. Sci. Paris Ser. D. 264,
1577 1579.
Ugolini F.C., Dahlgren R., La Manna J., Muhn W. &
Zachara J. (1991) Mineralogy and weathering
processes in Recent and Holocene tephra deposits
of the Pacific Northwest, USA. Geoderma, 51,
277 -299.
Vicente M.A., Razzaghe M. & Robert M. (1977)
Formation of aluminium hydroxy vermiculite (intergrade) and smectite from mica under acidic conditions. Clay Miner. 12, 101-112.
Vicente-Hemandez J., Vicente M.A., Robert M. &
Goodman B.A. (1983) Evolution des biotites en
fonction des conditions d'oxydo-reduction du milieu.
Clay Miner. 18, 267 275.
Villiers de J.M. & Jackson M.L. (1967) Aluminous
chlorite origin of pH dependent cation exchange
variations. Soil Sci. Soc. Am. Proc. 31,614-619.
Wada K. & Kakuto Y. (1983) lntegradient vermiculitekaolin mineral in a Korean Ultisol. Clays Clay
Miner. 31, 183-190.
Wada S.I. & Mizota C. (1982) Iron-rich halloysite with
crumpled lamellar morphology from Hokkaido,
Japan. Clays Clay Miner. 30, 315-317.
Walker G.F. (1949) The decomposition of biotite in the
soil. Mineral. Mag. 28, 693-703.
Watanabe T., Sowada Y., Russell J.D., McHardy W.J. &
Wilson M.J. (1992) The conversion of montmorillonite to interstratified halloysite-smectite by weathering in the Omi acid clay deposit, Japan. Clay
Miner. 27, 159-173.
Wilding L.P., Drees L.R., Smeck N.E. & Hall G.F.
(1971) Mineral and elemental composition of
Wisconsin age till deposits in west-central Ohio.
Pp. 290 318 in: Till: A Symposium (R.P.
Goldthwait, editor). Ohio State University Press,
Columbus, Ohio.
Wilke D.M. & Zech W. (1987) Mineralogies of silt and
clay fractions of twelve soil profiles in the Bolivian
Andes (Callavaya Region). Geoderma, 39, 189-208.

Origin and formation of clay minerals in soils"


Wilson M.J. (1965) Weathered biotite from Strathdon,
Aberdeenshire. Nature Lond. 210, 1188 1189.
Wilson M.J. (1966) Clay mineralogy of some soils
derived from a biotite-rich quartz gabbro in the
Strathdon area, Aberdeenshire. Clay Miner. 7,
91-100.
Wilson M.J. (1970) A study of weathering in a soil
derived from a b i o t i t e - h o r n b l e n d e rock. I.
Weathering of biotite. Clay Miner. 8, 291-303.
Wilson M.J. (1987) Soil smectites and related interstratified minerals: recent developments. Proc. lnt.
Clay Conf, Denver, 167-173.
Wilson M.J. (1993) Pedologic factors influencing the
distribution and properties of soil smectites. Trends
Agric. Sci. 1, 199-216.

25

Wilson M.J. & Mitchell B.D. (1979) Comparative study


of a Vertisol and an Entisol from the Blue Nile plains
of Sudan. Egypt J. Soil Sci. 19, 207-220.
Wilson M.J. & Nadeau P.H. (1985) Interstratified clay
minerals and weathering processes. In: The
Chemistry of Weathering (J.I. Drever, editor). D.
Riedel, Dordrecht, The Netherlands.
Wilson M.J., Bain D.C. & Duthie D.M.L. (1984) The
soil clays of Great Britain. II. Scotland. Clay Miner.
19, 709-735.
Wilson M.J., Oyegoke C. & Fraser A.R. (1997)
Occurrence of trioctahedral clay mica in some
Nigerian alluvial soils. 11th Int. Clay Conf.
Ottawa, Canada. Abstracts. p. A83.

You might also like